Subscriber access provided by University of Newcastle, Australia
Article
Molecular Mechanism of Dioxin Formation from Chlorophenol based on Electron Paramagnetic Resonance Spectroscopy Li Li Yang, Guorui Liu, Minghui Zheng, Yuyang Zhao, Rong Jin, Xiaolin Wu, and Yang Xu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b00828 • Publication Date (Web): 30 Mar 2017 Downloaded from http://pubs.acs.org on March 31, 2017
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 31
Environmental Science & Technology
1
Molecular Mechanism of Dioxin Formation from Chlorophenol
2
based on Electron Paramagnetic Resonance Spectroscopy
3
Lili Yang,†,‡, Guorui Liu,†,‡,*, Minghui Zheng,†,‡,*, Yuyang Zhao,†,‡2, Rong Jin,†,‡,
4
Xiaolin Wu,†,‡, Yang Xu,†,‡
5
†
6
Center for Eco-Environmental Sciences, Chinese Academy of Sciences, P.O. Box
7
2871, Beijing 100085, China
8
‡
9
*
State Key Laboratory of Environmental Chemistry and Ecotoxicology, Research
University of Chinese Academy of Sciences, Beijing 100049, China
Corresponding author.
10
E-mail:
[email protected] (Guorui Liu); Tel: 86 10 62849356; Fax: 86 10 62849355
11
[email protected] (Minghui Zheng); Tel: 86 10 62849172
12
1
ACS Paragon Plus Environment
Environmental Science & Technology
Page 2 of 31
13
ABSTRACT
14
Few studies have investigated the free radical intermediates involved in the formation
15
of
16
chlorophenol. This study clarified the reaction pathways during thermochemical
17
formation of PCDDs from 2,3,6-trichlorophenol (TCP) over a Cu(II)O/silica matrix,
18
which was used to simulate fly ash, at 298–523 K. The reaction was studied using
19
electron paramagnetic resonance (EPR) spectroscopy and theoretical calculations. In
20
situ EPR indicated the TCP radical (TCPR) formed by hydrogen abstraction of TCP.
21
Five elementary processes including dimerization of TCPR, ortho-chloride
22
abstraction, Smiles rearrangement, ring closure, and intra-annular elimination of Cl
23
were proposed to occur during formation of PCDDs. The proposed mechanism was
24
further confirmed by the detection of PCDD products from thermochemical
25
experiments
26
1,2,6,9-tetrachlorodibenzo-p-dioxin (TeCDD), 1,2,6,7-TeCDD, 1,2,8,9-TeCDD, and
27
1,4,6,9-TeCDD were detected by gas chromatography/quadrupole time-of-flight mass
28
spectrometry, and further confirmed by gas chromatography/high resolution mass
29
spectrometry. The detected PCDD products agree with the proposed PCDD formation
30
mechanism. Relatively high temperatures were found to lead to dechlorination of
31
TCPR to form phenoxy radicals in addition to PCDD/Fs. These radicals will be
32
attached to particles, which will increase their lifetimes. These reactions were further
33
verified by molecular orbital theory calculations. The discovery of persistent phenoxy
34
radicals is of environmental significance because of their potential toxicity. The
polychlorinated
in
a
dibenzo-p-dioxins
tube
furnace.
and
Several
dibenzofurans
dominant
2
ACS Paragon Plus Environment
(PCDD/Fs)
congeners,
from
including
Page 3 of 31
Environmental Science & Technology
35
details of this mechanism could be used for controlling PCDD/F formation during
36
industrial thermal processes.
37 38 39
INTRODUCTION
40
Polychlorinated dibenzo-p-dioxins and dibenzofurans (PCDD/Fs) have high
41
toxicity and bio-accumulate, can be unintentionally produced during multiple
42
industrial thermal processes.1 Currently, even though opinions diverge over the
43
formation mechanism of PCDD/Fs during thermal and catalytic processes, two
44
categories of formation pathways are widely recognized. The first category is
45
homogenous
46
chlorophenols and chlorobenzenes. The second category is heterogeneous reactions
47
involving de novo synthesis from carbon, chlorine, and oxygen, and chlorinated
48
organic precursors, which occurs under the catalysis of metallic oxides in fly ash.
49
Currently, PCDD/F formation pathways are mostly speculative, and are from studies
50
of homolog and congener patterns of PCDD/Fs in stack gas or fly ash collected from
51
full-scale facilities.2,3 However, the mechanistic details on the molecular level for
52
PCDD/F formation during thermochemical reactions, and especially the intermediates
53
involved in the reactions, have not been clarified to provide sufficient experimental
54
evidence for these reaction pathways.
reactions
involving
chlorinated
organic
precursors,
such
as
55
Chlorophenols are widely recognized to be important precursors for PCDD/Fs.
56
Chlorophenol levels are significantly correlated with unintentional production of 3
ACS Paragon Plus Environment
Environmental Science & Technology
57
PCDD/Fs from industrial thermal sources, and they can even be used as an indicator
58
for fast prediction of the potential PCDD/F emission levels in stack gas.4,5 Many
59
laboratory-scale studies have indicated that chlorophenols play a key role in PCDD/F
60
formation, and they are widely recognized as the predominant precursor under
61
pyrolytic or oxidative conditions.6 Chlorophenols are structurally similar to PCDD/Fs
62
and abundant in the stack gas from some thermal related industries.6,7 For instance,
63
PCDD/Fs can be formed in the oxidation of chlorinated phenols and chlorinated
64
benzenes under the catalysis of metal oxide particles or catalysts like manganese
65
peroxidase.6,8-10 In previous study, monochlorophenol and dichlorophenol have been
66
used to investigate the formation of PCDD/Fs.11,12 Trichlorophenol that easily produce
67
toxic tetra-chlorinated dioxin isomers during thermal reactions were scarcely studied.
68
In addition, based on molecular orbital theory calculations, trichlorophenol (TCP)
69
such as 2,3,6-TCP are believed to be important precursors for PCDD/F formation
70
because they have a high number of chlorine substitutes and relatively lower steric
71
hindrance because of oxygen–carbon coupling.13,14
72
PCDD/Fs can be formed in the cooling zone of industrial thermal processes,
73
which also generate large quantities of fly ash containing chlorinated organic
74
compounds and particle-supported metallic oxides.15 Under the catalytic effect of fly
75
ash containing metal compounds, PCDD/Fs can be easily produced during industrial
76
thermal processes.16 However, the complexity of the fly ash matrix poses challenges
77
for the detection of free radicals on the molecular scale. Consequently, a bottom-up
78
approach is used, in which the fly ash is separated into its organic, metal oxide, and 4
ACS Paragon Plus Environment
Page 4 of 31
Page 5 of 31
Environmental Science & Technology
79
particulate components, and the role of each of these components is assessed
80
individually and in combination with each other.17 It is widely recognized that copper
81
chlorides and oxides are the strongest catalysts for PCDD/F formation.18,19 Our
82
previous study on the occurrence of PCDD/Fs in industrial sources also indicated that
83
fly ash from secondary copper process had higher PCDD/F concentrations than that
84
from other secondary nonferrous smelting processes, primary nonferrous smelting
85
processes, and steel making processes.16 Many studies have investigated the oxidation
86
of chlorophenols and chlorobenzenes on the surface of silica-supported copper(II)
87
oxide and iron(III) oxide to simulate the fly ash substrate and clarify the PCDD/F
88
formation mechanism on such surfaces.9,20,21 Based on theoretical calculations, one
89
study proposed that condensation of two chlorophenoxy radicals is a significant
90
pathway for PCDD/Fs formation.22 Besides their contribution to PCDD/F formation,
91
organic free radicals such as chlorophenoxy radicals are potential toxic and persistent
92
in metal-containing particles, which suggested their environmental and health impacts
93
of thermal processes especially the copper-smelting industries.23
94
The details of the PCDD/F formation mechanism on the molecular level are
95
important for controlling PCDD/F formations in industrial thermal processes, and
96
identification of the intermediates, including organic free radicals, during the
97
thermochemical reactions of chlorophenol is required. In this study, 2,3,6-TCP was
98
used as a precursor for formation of PCDD through thermochemical reactions under
99
the catalysis of Cu(II)O-containing silica. This method was used to simulate the
100
particles–metallic oxide–chlorophenol system in fly ash. The occurrence of free 5
ACS Paragon Plus Environment
Environmental Science & Technology
101
radicals produced during the thermochemical reactions of 2,3,6-TCP were monitored
102
by electron paramagnetic resonance (EPR) spectroscopy, which could provide direct
103
evidence for understanding the mechanisms involved in PCDD formation. The
104
intermediates detected by EPR were used to propose a PCDD formation mechanism
105
using 2,3,6-TCP as the precursor, and this was verified by molecular orbital theory
106
calculations. The results provide important knowledge for controlling PCDD/F
107
formation during industrial thermal processes.
108 109
EXPERIMENTAL SECTION
110
In situ detection of free radicals involved in the thermochemical reactions of
111
2,3,6-TCP by EPR spectroscopy. EPR spectroscopy is a powerful tool for detecting
112
free radicals. In this study, the Dewar was inserted into the cavity of the EPR
113
spectrometer for in situ detection of free radicals during thermochemical reactions of
114
2,3,6-TCP. The kinetic behavior of radical intermediates produced during the thermal
115
reaction of chlorophenol were simulated and monitored. The Dewar was interfaced to
116
a nitrogen container to allow steady detection and identification of radicals generated
117
during the thermal process. Experiments were carried out by loading a 2,3,6-TCP
118
standard (> 98% pure, Tokyo Chemical Industry Co., Tokyo, Japan) or a mixture of
119
2,3,6-TCP and 5% Cu(II)O/SiO2 (Sinopharm Chemical Reagent Co., Shanghai, China)
120
into the EPR quartz tube (internal diameter (i.d.) = 4 mm, length = 10 mm) under a
121
constant nitrogen flow at a temperature between 298 K and 523 K. The silica particles
122
were heated in the air at 723 K for 4 h to remove potential organic contaminants on 6
ACS Paragon Plus Environment
Page 6 of 31
Page 7 of 31
Environmental Science & Technology
123
the surface. The quartz tube was located in an electrically heated Dewar under
124
atmospheric pressure for heating. After heating, the solid samples were cooled to 298
125
K and stored in the dark for 48 h. Then, the samples were analyzed again by EPR
126
spectroscopy to assess the influence of particulate matter (Cu(II)O/SiO2) on radical
127
lifetime.
128
All EPR spectra were recorded on a Bruker EMX-plus X-band EPR spectrometer
129
(Bruker Instruments, Billerica, MA). The instrument and operating parameters were
130
as follows: center field, 3520 G; microwave frequency, 9.36 GHz; microwave power,
131
0.63 mW; modulation frequency, 100 KHz; modulation amplitude, 1.0 G; sweep
132
width, 200 G; receiver gain, 30 dB; and time constant, 0.01 ms. Radical quantification
133
was conducted using Bruker’s Xenon program according to the quantitative theory of
134
spin calculation (see the Supporting Information).
135
Thermochemical experiments for PCDD/F formation from 2,3,6-TCP in a
136
tube furnace for capturing the reaction products. Thermochemical experiments
137
were conducted to simulate the formation conditions of radicals and PCDD/Fs during
138
industrial thermal process. Thermochemical experiments were performed using a tube
139
furnace (GSL-1100X, Kejing Material Technology Co., Hefei, China) equipped with
140
a quartz tube. A diagram of the apparatus is shown in Figure S1. Air was passed
141
through the tube furnace at a constant flow rate of 50 mL min−1. Silica is the main
142
component of fly ash (5−50%), and always acts as the support for transition metals in
143
catalytic systems.8 As a surrogate for these particles, we used Cu(II)O (5% mass
144
fraction) supported on silica. This was mixed with 0.1 g of 2,3,6-TCP standard to 7
ACS Paragon Plus Environment
Environmental Science & Technology
145
create a system to study interactions among specific, well-defined components of the
146
fly ash. The reactant was placed in a porcelain boat, which was then placed in the
147
middle of the furnace, and the furnace was heated to 250 °C for 15 min for the in situ
148
reaction. The furnace temperature was increased at a rate of 10 °C min−1. Based on
149
the results of our previous study, more than 90% of the PCDD/Fs were in the gas
150
phase at a reaction temperature of 523 K.24 The gaseous products of the thermal
151
process were captured by a toluene-filled absorption bottle held in an ice bath and
152
connected to the end of the reaction tube. After heating, the reactor (a quartz tube) and
153
all of the fittings were rinsed with toluene to recover any gas products deposited on
154
the surfaces of the tube. The rinsing solution was then combined with the toluene
155
from the absorption bottles. The solid residues and gas phase samples were extracted
156
and cleaned according to a modification of the method described in our previous
157
study.24
158
In addition, blank and contrast experiments were also carried out. The blank
159
experiments involved running the tube furnace containing only SiO2, and again
160
containing all reactants except the chlorophenols. The contrast experiment used a
161
chlorophenol mixture without Cu(II)O. The blank samples were used to determine the
162
quantity of PCDD in the thermal reaction products generated from the particle
163
surrogate samples, and this value was compared with that from the final samples. The
164
contrast sample was used to evaluate the influence of metallic oxide on the formation
165
of PCDDs. The yields of the products were calculated for comparison of the contrast
166
samples with the experimental samples using the following equation:25,26 8
ACS Paragon Plus Environment
Page 8 of 31
Page 9 of 31
Environmental Science & Technology
[ௗ௨௧]
ܻ݈݅݁݀ = (
167
[ோ௧௧]
) × 100%,
168
where [Product] is the concentration of PCDDs formed; [Reactant] is the initial
169
concentration of the reactants injected into the reactor tube; and A is the molar
170
stoichiometric factor, which was identified as two in this case because two reactant
171
molecules (2,3,6-TCP) can form one PCDD molecule.
172
Determination of the reaction products by gas chromatography quadrupole
173
time-of-flight
tandem
mass
spectrometry
and
174
chromatography with high-resolution mass spectrometry. The gaseous products
175
containing more than 90% of the total PCDD/Fs formed in the thermal process were
176
cleaned using a multilayer silica gel column treated with 44% (mass fraction) sulfuric
177
acid and 33% (mass fraction) sodium hydroxide, concentrated to a volume of about 20
178
µL using a rotary evaporator and a gentle stream of N2, and detected by gas
179
chromatography combined with a quadrupole time-of-flight mass spectrometry (GC
180
Q-TOF MS) (Agilent Technologies, Santa Clara, CA) and high-resolution gas
181
chromatography/high-resolution mass spectrometry (HRGC/HRMS). The GC Q-TOF
182
MS measurements were carried out on an Agilent 7890 GC system coupled to a TOF
183
mass spectrometer, operating in electron ionization mode. The GC separation was
184
performed using a DB-5 MS fused silica capillary column (30 m × 0.25 mm i.d., 0.25
185
µm film thickness; Agilent Technologies). The oven temperature was programmed as
186
follows: 80 °C (hold 1 min); increased at 20 °C/min to 120 °C; increased at 8 °C/min
187
to 200 °C; increased at 20 °C/min to 250 °C; increased at 5 °C/min to 310 °C (hold 15
188
min). Split injections of 1 µL of sample were carried out. Helium was used as carrier 9
ACS Paragon Plus Environment
high-resolution
gas
Environmental Science & Technology
189
gas at 1 mL/min. The interface and source temperatures were set to 300 °C and
190
250 °C, respectively, and a solvent delay of 4 min was selected. In the TOF MS, five
191
spectra were acquired per second in the mass range m/z 50–650. The qualifications
192
were performed by NIST spectrum library matching. “Unknown compounds analysis”
193
software in the Mass Hunter Workstation (Agilent Technologies) was also used for
194
purification and classification of co-outflow peaks by deconvolution of the data sets,
195
which could improve the unknown compounds matching accuracy with the spectral
196
library and established an error limit of 2.3 ppm for identification. Accurate
197
identification and quantification of PCDD/Fs in the gas phase were performed by
198
HRGC/HRMS using a DB-5ms capillary column (60 m × 0.25 mm i.d., 0.25 µm film
199
thickness; Agilent Technologies). The HRMS was operated in selected-ion monitoring
200
mode at a resolution of ≥ 10,000. Details for the HRGC/HRMS procedure were
201
described in the supporting information.27
202
Simulation of EPR spectrum and density functional theory calculation used
203
for mechanism verification. The Gaussian 09 suite of programs was used for density
204
functional theory analysis of electronic structure calculations.28 MPWB1K is a hybrid
205
functionals that is based on the modified Perdew and Wang exchange functional and
206
Becke’s 1995 correlation functional, and has been successfully used for the prediction
207
of transition state geometries and thermochemical kinetics.14 Therefore, as a
208
reasonable compromise between computation time and accuracy, the MPWB1K
209
method and a standard 6-31+G(d,p) basis set were chosen for optimization of the
210
geometries of the stationary points.29,30 Vibrational frequencies were calculated at the 10
ACS Paragon Plus Environment
Page 10 of 31
Page 11 of 31
Environmental Science & Technology
211
same level to check each geometry of the potential energy surface minimum.
212 213
RESULTS AND DISCUSSION
214
EPR detection and kinetic calculations of free radicals involved in the
215
thermochemical reactions of 2,3,6-TCP. The most direct route for PCDD/F
216
formation is the gas-phase reaction of chemical precursors.31 Therefore, 2,3,6-TCP
217
was added on its own to the EPR quartz tube for in situ EPR spectroscopy as the
218
temperature was increased from 298 K to 523 K. The results were used to monitor the
219
generation of free radicals for comparison with the fly ash simulation system. A large
220
quantity of free radicals was generated when the temperature increased (Fig. 1), and
221
the concentrations of free radicals were 1.906 × 1017 spins/mm3 and 1.559 × 1017
222
spins/mm3 at 423 K and 523 K, respectively. However, when the temperature dropped
223
to 298 K, the concentration of free radicals detected was twice that at 523 K, which
224
indicated that free radicals were generated in the same zone as PCDD/Fs formed in,
225
which is the cooling zone in the industrial thermal processes. When placed at room
226
temperature (298 K) for 48 h, an EPR signal was not detected.
227
It has shown that dimerization of chlorophenoxy radicals is the dominant
228
pathway in the gas-phase formation of PCDD/Fs from chlorophenol precursors.32
229
Thus, the formation of chlorophenoxy radicals is the initial step in the formation of
230
PCDD/Fs. Besides, only chlorophenols with chlorine at the ortho position are
231
considered capable of forming PCDDs, while para-chlorophenol is responsible for
232
PCDF formation.33 Therefore, hydrogen (H) abstraction and dechlorination reaction 11
ACS Paragon Plus Environment
Environmental Science & Technology
233
process were calculated accurately to assess if formation of chlorophenoxy radicals
234
and PCDD/Fs by these pathways is reasonable. The potential barriers (∆E) and the
235
heats of reaction (∆H), which both include the ZPE correction, were calculated at the
236
MPWB1K/6-31+G(d,p) level. Calculated relative energies are given in Figure 2, and
237
optimized geometries of H abstraction and dechlorinated transition states are denoted
238
as TS1–TS21 in Figure S2. The H abstraction reaction can occur readily with
239
unimolecular, bimolecular, or possibly other low-energy pathways like heterogeneous
240
reactions.34,35 The unimolecular reaction occurs via the cleavage of the O-H bond of
241
chlorophenol. The bimolecular reactions proceed through the phenoxyl-H abstraction
242
by active and abundant radicals like H, OH, and Cl in the combustion system. H
243
abstraction occurs through a bimolecular reaction of the H radical, and is barrier-less
244
as the transition states with the ZPE correction are lower than the energies of reactants
245
(Figure 2). In addition, the process is exothermic, and is an energetically feasible
246
reaction for the formation of the TCP radical (TCPR) under pyrolysis or combustion
247
conditions. Afterwards, the EPR spectrum of the phenoxy radical suggests that
248
hydrodechlorination of the chlorophenoxy radical can occur along with the increasing
249
temperature. The dechlorination process has a relatively high heat of reaction and is
250
endothermic, whereas the hydrogenation process is exothermic and can easily occur
251
with no barrier to electron pairing (Figure 2). Therefore, it is easy to convert
252
2,3,6-TCP to 2,3,6-TRPR by H abstraction, and phenoxy radicals can be produced by
253
hydrodechlorination only if there is sufficient thermal energy, which is consistent with
254
our experimental results. 12
ACS Paragon Plus Environment
Page 12 of 31
Page 13 of 31
Environmental Science & Technology
255
Persistent free radical formation of 2,3,6-TCP on the surface of
256
silica-supported Cu(II) oxide. Cu(II)O-containing silica was used as the thermal
257
reaction substrate to simulate the fly ash system in this study. 2,3,6-TCP and 5%
258
Cu(II)O/SiO2 were mixed thoroughly and placed in the EPR quartz tube to a height of
259
1 cm for the in situ detection. The mixture of 2,3,6-TCP and 5% Cu(II)O/SiO2 did not
260
generate free radicals at room temperature (Figure 3). When the temperature was
261
increased to 423 K or 523 K, strong EPR signals were detected, and the free radical
262
concentration was 6.158 × 1017 spins/mm3, which was about 38 times that of the same
263
quantity of 2,3,6-TCP at the same temperature. The large increase in the radical
264
concentration indicated that the Cu(II)O/SiO2 could trigger free radical formation.
265
When the temperature dropped to 298 K, the concentration of free radicals was 4.427
266
× 1018 spins/mm , and seven times that at 523 K. This result further implied that free
267
radical and PCDD/F formation occur in the cooling zones of thermal processes. When
268
placed at room temperature (298 K) for 48 h, a strong EPR signal quantified as 2.187
269
× 1018 spins/mm was observed when the particulate matter and metallic oxide were
270
present, which was different to the result obtained for the reaction of 2,3,6-TCP on its
271
own. This result suggested that metal particles could extend the lifetime of free
272
radicals produced by chlorophenol. Furthermore, the g value of the free radicals
273
shifted from 2.0042 at 423 K to 2.0035 at 298 K after heating, which suggests that
274
type of free radical formed may change from oxygen-centered to a more stabilized
275
carbon-centered phenoxy radical.36
276
3
3
Figure 4 shows the dynamic changes in free radical intermediates formed in the 13
ACS Paragon Plus Environment
Environmental Science & Technology
277
particulate–metallic oxide–chlorophenol system when the temperature was increased
278
from 298 K to 523 K. When at the temperature was around 350 K, hyperfine splitting
279
of ten peaks was detected. The results of simulation by the Isotropic Radicals program
280
inferred that the hyperfine signals were generated by interaction of three chlorine
281
nuclear spins and one electron spin of the TCPR. In general, a temperature increase
282
will trigger H abstraction of TCP to form TCPR at around 350 K, and further
283
increases in the temperature will lead to dechlorination of TCPR to form phenoxy
284
radicals. This indicated that the non-chlorine substituted phenoxyl-radical was much
285
more stable than that of chlorine substituted phenoxy-radicals during the thermal
286
processes, which was in agreement with that reported by Vejerano et al.37 As expected,
287
the phenoxy radicals formed can persist and are stabilized by chemisorption to the
288
metal oxide, with a single electron transfer to the metal cation center.17,38
289
Dela Cruz et al. proposed three possible pathways for environmentally persistent
290
free radical (EPFR) formation from pentachlorophenol on the surface of Fe(III) oxide,
291
which can also be adapted to the formation of the EPFRs of TCP on the surface of Cu
292
(II) oxide (Figure 5).39 When chlorophenol is first absorbed on the Cu(II) oxide
293
surface by physical absorption, loss of H2O occurs to form phenolate with
294
simultaneous or rapid sequential electron transfer from the phenolate to Cu(II), which
295
produces Cu(I) and the TCP EPFR. Oxygen-centered and carbon-centered radicals
296
can transform from one to the other. The second pathway is chemisorption through
297
loss of HCl or both HCl and H2O, and formation of a bichlorosemiquinone radical.
298
Only the first and second pathways were observed in soils contaminated with 14
ACS Paragon Plus Environment
Page 14 of 31
Page 15 of 31
Environmental Science & Technology
299
chlorophenols under cool-zone conditions at temperatures between 150 and 600 °C,
300
even though the possibility of the third pathway cannot be ruled out in soil at ambient
301
temperatures with reaction times of years.39 However, using the in situ reaction, we
302
successfully monitored the hyperfine splitting signals of TCPR, which is the product
303
of the first pathway. The result suggested that this pathway was the dominant
304
formation mechanism of EPFRs during industrial thermal reactions at temperatures
305
between 298 K and 423 K. In this temperature range, persistent phenoxy radicals
306
could be formed along with PCDD/Fs in heterogeneous synthesis.
307
PCDD formation of TCPR during thermochemical reactions in a tube
308
furnace. The tube furnace reactions for simulation of industrial thermal process were
309
first analyzed by GC Q-TOF MS as described in the experimental section. Figure S3
310
shows the total ion chromatograms obtained by analysis of the gaseous products in
311
positive ionization modes. Peaks for 2,3,6-TCP and other chlorophenols eluted before
312
17 min. The principle products, PCDDs, PCDFs, polybrominated diphenyl ether, and
313
polychloro-3-(polychlorophenoxy)-dibenzo dioxin, were eluted after 17 min. In total,
314
29 compounds were tentatively identified (Table 1) in the gas phase using TOF MS
315
data
316
tetrachlorodibenzo-p-dioxins (TeCDDs) (Figure S3) were the main products and gave
317
the highest responses, which were approximately 102 times than those of PeCDDs and
318
HxCDDs, 103 times those of HpCDDs, and 104 times those of OCDD. The responses
319
of
320
polychloro-3-(polychlorophenoxy)-dibenzo dioxin were even lower. Formation of
in
high
PCDFs,
resolution
mode
with
polychlorinated
an
error
below
diphenyl
15
ACS Paragon Plus Environment
2.3
ppm.
ether,
Four
and
Environmental Science & Technology
321
PCDD/Fs by the thermochemical reactions of 2-chlorophenol were previously
322
reported.12 Mono- and di-chlorinated congeners were expected to be the dominant
323
PCDD/Fs. Actually, several products including 1-monochlorodibenzo-p-dioxin
324
(1-MCDD), dibenzofuran (DF), chloronaphthalene were detected by GC/MS when
325
using 2-chlorophenol as precursor.12 In this study, four tetrachlorodibenzo-p-dioxins
326
(TeCDDs) isomers were the dominant products formed by thermochemical reactions
327
of 2,3,6-TCP. Although penta- to octa- chlorinated dioxin congeners were also
328
detected by GC-TOF MS, their contents were far lower than that of TeCDD isomers.
329
It is widely recognized that PCDDs are formed via the o-phenoxy-phenol
330
intermediate through oxygen–carbon coupling of TCPRs.13 Because of the
331
asymmetric chlorine substitution, three o-phenoxy-phenol intermediates (IM1-IM3 in
332
Figure 6) can be formed by TCPR coupling. We propose the formation of TeCDDs in
333
gas phase radical–radical reactions (Figure 6) involves five elementary processes as
334
follows: dimerization of TCPRs, ortho-chloride abstraction, Smiles rearrangement,
335
ring closure, and intra-annular elimination of Cl. These five elementary reactions lead
336
to 1,2,6,9-TeCDD, 1,2,6,7-TeCDD, 1,4,6,9-TeCDD, and 1,2,8,9-TeCDD as the main
337
products `from TCPR.
338
To identify and quantify the products of the tube furnace reaction, and to verify
339
our proposed PCDD formation mechanism, isotope dilution HRGC/HRMS was used
340
for further determination of the experimental samples and contrast samples in the tube
341
furnace simulation experiment. Four TeCDDs (Figure S4) were identified by
342
comparing the retention times with those of corresponding native standard compounds. 16
ACS Paragon Plus Environment
Page 16 of 31
Page 17 of 31
Environmental Science & Technology
343
The main products for the chromatographic peaks at 23.20 min, 24.90 min, 26.43 min,
344
and 27.20 min were identified as 1,4,6,9-TeCDD, 1,2,6,9-TeCDD, 1,2,6,7-TeCDD,
345
and 1,2,8,9-TeCDD, respectively, which were the same as our expected products in
346
gas phase reactions. The yields of 1,2,6,9-TeCDD (5.7%, percentage in the bracket
347
indicated the yields) and 1,2,6,7-TeCDD (3.3%) were higher than those of
348
1,4,6,9-TeCDD (0.3%) and 1,2,8,9-TeCDD (2.4%) in the experimental samples,
349
which is in accordance with the speculation in Figure 6 that more pathways lead to
350
formation of 1,2,6,9-TeCDD and 1,2,6,7-TeCDD than 1,4,6,9-TeCDD and
351
1,2,8,9-TeCDD. The consistency between the experimental observations and our
352
speculations for PCDD formation indicates the proposed mechanism is reasonable and
353
valid. In addition, blank samples containing all other reactants except the
354
chlorophenols were used to confirm that there was no obvious contamination from the
355
Cu(II)O/SiO2 reaction substrate, whose product yield was less than 0.0017% of that
356
obtained from the experimental samples. The contrast sample containing the
357
chlorophenol mixture without Cu(II)O was also used to investigate the catalytic
358
activity of metallic oxide for PCDD formation in the thermal process. The yields of
359
1,4,6,9-TeCDD, 1,2,6,9-TeCDD, 1,2,6,7-TeCDD, and 1,2,8,9-TeCDD for the contrast
360
sample without the catalysis of Cu(II)O were 0.005%, 0.014%, 0.010%, and 0.006%,
361
respectively. These results were all less than the 1.5% of those achieved for the
362
experimental samples containing 5% Cu(II)O/SiO2. Furthermore, the gas phase
363
reactions in Figure 6 only occurred at temperatures above 873 K, and the PCDD
364
formation results were the same as those in our simulation at 523 K, which indicates 17
ACS Paragon Plus Environment
Environmental Science & Technology
365
that this is a surface-mediated process.40 This result implied that metallic oxides such
366
as Cu(II)O are important for both stabilization of free radicals through chemisorption
367
and electron transfer and catalysis of PCDD formation from chlorophenol.
368 369
Supporting Information
370
Details of HRGC/HRMS and radical quantification method, Figure S1-S4. This
371
information is available free of charge via the Internet at http://pubs.acs.org/.
372 373
AUTHOR INFORMATION
374
Notes
375
The authors declare no competing financial interest.
376 377
ACKNOWLEDGMENTS
378
This work was supported by the Chinese National 973 Program (Grant No.
379
2015CB453100), the National Natural Science Foundation of China (Grant No.
380
91543108), the Strategic Priority Research Program of the Chinese Academy of
381
Sciences (Grant No. XDB14020102), and the Youth Innovation Promotion
382
Association of the Chinese Academy of Sciences (Grant No. 2016038). We appreciate
383
Dr. Jiajia Wu from Agilent Technologies for her assistance on the GC-TOF MS
384
analysis.
385 386
REFERENCE: 18
ACS Paragon Plus Environment
Page 18 of 31
Page 19 of 31
Environmental Science & Technology
387 388 389 390 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430
(1) Liu, G. R.; Jiang, X. X.; Wang, M.; Dong, S. J.; Zheng, M. H. Comparison of PCDD/F levels and profiles in fly ash samples from multiple industrial thermal sources. Chemosphere 2015, 133, 68-74. (2) Stanmore, B. R. Modeling the formation of PCDD/F in solid waste incinerators. Chemosphere 2002, 47, 565-573. (3) Khachatryan, L.; Burcat, A.; Dellinger, B. An elementary reaction-kinetic model for the gas-phase formation of 1,3,6,8- and 1,3,7,9-tetrachlorinated dibenzo-p-dioxins from 2,4,6–trichlorophenol. Combustion and Flame 2003, 132, 406-421. (4) Lavric, E.; Konnov, A.; Ruyck, J. Modeling the formation of precursors of dioxins during combustion of woody fuel volatiles. Fuel 2005, 84, 323-334. (5) Blumenstock, M.; Zimmermann, R.; Schramm, K. W.; Kaune, A.; Nikolai, U.; Lenoir, D.; Kettrup, A. Estimation of the dioxin emission (PCDD/FI-TEQ) from the concentration of low chlorinated aromatic compounds in the flue and stack gas of a hazardous waste incinerator. J. Anal. Appl. Pyrolysis 1999, 49, 179-190. (6) Fernandez-Castro, P.; San Roman, M. F.; Ortiz, I. Theoretical and experimental formation of low chlorinated dibenzo-p-dioxins and dibenzofurans in the Fenton oxidation of chlorophenol solutions. Chemosphere 2016, 161, 136-144. (7) Ryu, J. Y.; Mulholland, J. A.; Kim, D. H.; Takeuchi, M. Homologue and isomer patterns of polychlorinated dibenzo-p-dioxins and dibenzofurans from phenol precursors: comparison with municipal waste incinerator data. Environ Sci Technol 2005, 39, 4398-4406. (8) Mosallanejad, S.; Dlugogorski, B. Z.; Kennedy, E. M.; Stockenhuber, M.; Lomnicki, S. M.; Assaf, N. W.; Altarawneh, M. Formation of PCDD/Fs in Oxidation of 2-Chlorophenol on Neat Silica Surface. Environ Sci Technol 2016, 50, 1412-1418. (9) Nganai, S.; Dellinger, B.; Lomnicki, S. PCDD/PCDF ratio in the precursor formation model over CuO surface. Environ Sci Technol 2014, 48, 13864-13870. (10) Munoz, M.; Gomez-Rico, M. F.; Font, R. PCDD/F formation from chlorophenols by lignin and manganese peroxidases. Chemosphere 2014, 110, 129-135. (11) Weber, R.; Hagenmaier, H. Mechanism of the formation of polychlorinated dibenzo-p-dioxins and dibenzofurans from chlorophenols in gas phase reactions. Chemosphere 1999, 38, 529-549. (12) Evans, C. S.; Dellinger, B. Mechanisms of dioxin formation from the high-temperature pyrolysis of 2-chlorophenol. Environ. Sci. Technol. 2003, 37, 1325-1330. (13) Zhang, Q.; Li, S.; Qu, X.; Shi, X.; Wang, W. A quantum mechanical study on the formation of PCDD/Fs from 2-chlorophenol as precursor. Environ Sci Technol 2008, 42, 7301-7308. (14) Zhang, Q.; Yu, W.; Zhang, R.; Zhou, Q.; Gao, R.; Wang, W. Quantum chemical and kinetic study on dioxin formation from the 2,4,6-TCP and 2,4-DCP precursors. Environ Sci Technol 2010, 44, 3395-3403. (15) Altwicker, E. R.; Konduri, R. K. N. V.; Lin, C.; Milligan, M. S. Rapid Formation of Polychlorinated Dioxins Furans in the Post Combustion Region during Heterogeneous Combustion. Chemosphere 1992, 25, 1935-1944. (16) Liu, G.; Jiang, X.; Wang, M.; Dong, S.; Zheng, M. Comparison of PCDD/F levels and profiles in fly ash samples from multiple industrial thermal sources. Chemosphere 2015, 133, 68-74. (17) Nwosu, U. G.; Roy, A.; dela Cruz, A. L.; Dellinger, B.; Cook, R. Formation of environmentally persistent free radical (EPFR) in iron(III) cation-exchanged smectite clay. Environmental Science: Processes & impacts 2016, 18, 42-50. (18) Chang, S. H.; Yeh, J. W.; Chein, H. M.; Hsu, L. Y.; Chi, K. H.; Chang, M. B. PCDD/F adsorption and destruction in the flue gas streams of MWI and MSP via Cu and Fe catalysts supported on carbon. 19
ACS Paragon Plus Environment
Environmental Science & Technology
431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469 470 471 472 473 474
Page 20 of 31
Environ Sci Technol 2008, 42, 5727-5733. (19) Chin, Y. T.; Lin, C.; Chang-Chien, G. P.; Wang, Y. M. PCDD/F Formation Catalyzed by the Metal Chlorides and Chlorinated Aromatic Compounds in Fly Ash. Aerosol Air Qual Res 2012, 12, 228-236. (20) Caserini, S.; Monguzzi, A. M. PCDD/Fs emissions inventory in the Lombardy Region: results and uncertainties. Chemosphere 2002, 48, 779-786. (21) Lomnicki, S.; Dellinger, B. A detailed mechanism of the surface-mediated formation of PCDD/F from the oxidation of 2-chlorophenol on a CuO/silica surface. J Phys Chem A 2003, 107, 4387-4395. (22) Louw, R.; Ahonkhai, S. I. Radical/radical vs radical/molecule reactions in the formation of PCDD/Fs from (chloro)phenols in incinerators. Chemosphere 2002, 46, 1273-1278. (23) Vejerano, E.; Lomnicki, S.; Dellinger, B. Formation and Stabilization of Combustion-Generated Environmentally Persistent Free Radicals on an Fe(III)(2)O-3/Silica Surface. Environ Sci Technol 2011, 45, 589-594. (24) Wang, M.; Liu, G.; Jiang, X.; Xiao, K.; Zheng, M. Formation and potential mechanisms of polychlorinated dibenzo-p-dioxins and dibenzofurans on fly ash from a secondary copper smelting process. Environmental science and Pollution Research 2015, 22, 8747-8755. (25) Nganai, S.; Lomnicki, S.; Dellinger, B. Ferric oxide mediated formation of PCDD/Fs from 2-monochlorophenol. Environ Sci Technol 2009, 43, 368-373. (26) Nganai, S.; Lomnicki, S.; Dellinger, B. Formation
of
PCDD/Fs from oxidation
of
2-monochlorophenol over an Fe2O3/silica surface. Chemosphere 2012, 88, 371-376. (27) Yang, L.; Liu, G.; Zheng, M.; Jin, R.; Zhu, Q.; Zhao, Y.; Zhang, X.; Xu, Y. Atmospheric occurrence and health risks of PCDD/Fs, polychlorinated biphenyls, and polychlorinated naphthalenes by air inhalation in metallurgical plants. Sci. Total Environ. 2017, 580, 1146-1154. (28) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Gill, P. W. M.; Johnson, B. G.; Robb, M. A.; Cheeseman, J. R.; Keith, T. A.; Petersson, G. A.; Montgomery, J. A.; Raghavachari, K.; Allaham, M. A.; Zakrzewski, V. G.; Ortiz, J. V.; Foresman, J. B.; Cioslowski, J.; Stefanov, B. B.; Nanayakkara, A.; Challacombe, M.; Peng, C. Y.; Ayala, P. Y.; Chen, W.; Wong, M. W.; Andres, J. L.; Replogle, E. S.; Gomperts, R.; Martin, R. L.; Fox, D. J.; Binkley, J. S.; Defrees, D. J.; Baker, J.; Stewart, J. P.; Head-Gordon, M.; Gonzales, C.; Pople, J. A. Gaussion, INC., Wallingford CT. 2010. (29) Zhao, Y.; Truhlar, D. G. Hybrid meta density functional theory methods for thermochemistry, thermochemical kinetics, and noncovalent interactions: The MPW1B95 and MPWB1K models and comparative assessments for hydrogen bonding and van der Waals interactions. J Phys Chem A 2004, 108, 6908-6918. (30) Liu, G.; Zhan, J.; Zhao, Y.; Li, L.; Jiang, X.; Fu, J.; Li, C.; Zheng, M. Distributions, profiles and formation mechanisms of polychlorinated naphthalenes in cement kilns co-processing municipal waste incinerator fly ash. Chemosphere 2016, 155, 348-357. (31) Khachatryan, L.; Asatryan, R.; Dellinger, B. Development of expanded and core kinetic models for the gas phase formation of dioxins from chlorinated phenols. Chemosphere 2003, 52, 695-708. (32) Evans, C. S.; Dellinger, B. Mechanisms of dioxin formation from the high-temperature oxidation of 2-chlorophenol. Environ Sci Technol 2005, 39, 122-127. (33) Weber, R.; Hagenmaier, H. Mechanism of the formation of polychlorinated dibenzo-p-dioxins and dibenzofurans from chlorophenols in gas phase reactions. Chemosphere 1999, 38, 529-549. (34) Evans, C. S.; Dellinger, B. Mechanisms of dioxin formation from the high-temperature pyrolysis of 2-bromophenol. Environ Sci Technol 2003, 37, 5574-5580. (35) Altarawneh, M.; Dlugogorski, B. Z.; Kennedy, E. M.; Mackie, J. C. Quantum chemical and kinetic 20
ACS Paragon Plus Environment
Page 21 of 31
Environmental Science & Technology
475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490
study of formation of 2-chlorophenoxy radical from 2-chlorophenol: unimolecular decomposition and bimolecular reactions with H, OH, Cl, and O2. J Phys Chem A 2008, 112, 3680-3692. (36) Dellinger, B.; Lomnicki, S.; Khachatryan, L.; Maskos, Z.; Hall, R. W.; Adounkpe, J.; McFerrin, C.; Truong, H. Formation and stabilization of persistent free radicals. Proceedings of the Combustion Institute. International Symposium on Combustion 2007, 31, 521-528. (37) Vejerano, E.; Lomnicki, S. M.; Dellinger, B. Formation and Stabilization of Combustion-Generated, Environmentally Persistent Radicals on Ni(II)O Supported on a Silica Surface. Environ. Sci. Technol. 2012, 46, 9406-9411. (38) Kiruri, L. W.; Khachatryan, L.; Dellinger, B.; Lomnicki, S. Effect of copper oxide concentration on the formation and persistency of environmentally persistent free radicals (EPFRs) in particulates. Environ Sci Technol 2014, 48, 2212-2217. (39) Dela Cruz, A. L. N.; Gehling, W.; Lomnicki, S.; Cook, R.; Dellinger, B. Detection of Environmentally Persistent Free Radicals at a Superfund Wood Treating Site. Environ Sci Technol 2011, 45, 6356-6365. (40) Chatterjee, D.; Deutschmann, O.; Warnatz, J. Detailed surface reaction mechanism in a three-way catalyst. Faraday Discuss. 2001, 119, 371-384.
21
ACS Paragon Plus Environment
Environmental Science & Technology
Page 22 of 31
491
Table and Figure Legends:
492
Table 1. Compounds tentatively identified in the gas phase from the tube furnace
493
reaction, and the main parameters that support their identification
494
Figure 1. Electron spin resonance spectra of 2,3,6-trichlorophenol at (a) 298 K, (b)
495
423 K, (c) 523 K, and (d) 298 K after heating and (e) 48 h after heating.
496
Figure 2. Schematic of the formation of 2,3,6-trichlorophenol radical and phenoxy
497
radical from the reactions of 2,3,6-trichlorophenol with H. Potential barriers (∆E,
498
kcal/mol) and the heats of reaction (∆H, kcal/mol) are shown.
499
Figure 3. Electron spin resonance spectra of 2,3,6-trichlorophenol on the surface of
500
silica-supported copper(II) oxide at (a) 298 K, (b) 423 K, (c) 523 K, and (d) 298 K
501
after heating and (e) 48 h after heating.
502
Figure 4. Dynamic changes in the electron spin resonance spectra of
503
2,3,6-trichlorophenol on the surface of silica-supported copper(II) oxide at 298 K to
504
523 K. The inset shows a comparison of the experimental (green) and the simulated
505
(red) spectra.
506
Figure 5. Three possible pathways for the formation of persistent free radicals of
507
2,3,6-trichlorophenol radical from
508
silica-supported copper(II) oxide.39
509
Figure 6. Proposed formation routes of tetrachlorodibenzo-p-dioxins (TeCDDs)
510
involving the following five elementary processes: dimerization of 2,3,6-
511
trichlorophenol radicals, ortho-chloride abstraction, Smiles rearrangement, ring
512
closure, and intra-annular elimination of chloride.
2,3,6-trichlorophenol on
22
ACS Paragon Plus Environment
the
surface
of
Page 23 of 31
Environmental Science & Technology
Table 1. Compounds tentatively identified in the gas phase from the tube furnace reaction, and the main parameters that support their identification. Name
TeCDDs
Rention time
GC-QTOF Ion (m/z)
Formula
Error (ppm)
17.56
-0.5
17.88
-0.7
18.17
319.896
C12H4Cl4O2
18.31
Proposed structure Cl
Cl O
-0.7
O Cl
-0.2
Cl Cl
Cl
Polychlorinat ed diphenyl ether
O
Cl
18.02 Cl
303.9011
C12H4Cl6O
Cl Cl
--
Cl
Cl O
18.91 Cl
18.86 PeCDF
-1.4 337.8622
C12H3Cl5O
19.55
0
19.02
-1.1
19.4
-0.9
Cl
Cl Cl
Cl Cl
O
Cl
Cl
Cl Cl
Cl O
PeCDDs
353.8571
C12H3Cl5O2
19.75
Cl
O
-0.7 Cl
Polychlorinat ed diphenyl ether
HxCDDs
Cl Cl
C12H3Cl7O
19.89
Cl
Cl
Cl
--
407.7998 20.31
Cl O
C12H3Cl7O
Cl Cl
20.63
-0.4
21.07
-0.4
21.13 21.49
387.818
C12H2Cl6O2
-0.4 -0.5
21.59
-0.5
21.65
-0.6
20.44
0
HxCDFs
371.8232 20.92
0
22.94
-2.3
HpCDDs
421.7791
Cl
O
Cl
O
Cl
Cl
Cl Cl
C12H2Cl6O
Cl
Cl
Cl O
Cl
Cl
Cl
Cl
Cl O
Cl
C12HCl7O2
23.45
0
Cl
O Cl
22.73
Cl Cl
Cl
Cl O
HpCDF
405.7842
C12HCl7O
Cl
0 Cl
Cl Cl
23.22 23
ACS Paragon Plus Environment
Cl
Environmental Science & Technology
Page 24 of 31
Cl
Cl
Cl
OCDD
25.45
455.7401
C12Cl8O2
O
Cl
-0.7 O
Cl
Cl
Cl
Polychlorinat ed diphenyl ether Tetrachloro-3 -(dichloro-ph enoxy)-diben zo dioxin Pentachloro3-(dichloro-p henoxy)-dibe nzo dioxin
Cl
Cl
Cl
Cl
21.26
441.7614
C12H2Cl8O
Cl
O
-Cl
Cl
Cl Cl
Cl
Cl O
27.58
479.8456
C18H6Cl6O3
O
-O
Cl Cl
Cl
Cl
Cl
Cl O
32.64
513.8058
C18H5Cl7O3
Cl
O
-Cl Cl
24
ACS Paragon Plus Environment
O Cl
Cl
Page 25 of 31
Environmental Science & Technology
48 h after heating
298 K after heating 523 K 423 K 298 K 3350
3360
3370
3380
Figure 1. Electron spin resonance spectra of 2,3,6-trichlorophenol at (a) 298 K, (b) 423 K, (c) 523 K, and (d) 298 K after heating and (e) 48 h after heating.
25
ACS Paragon Plus Environment
Environmental Science & Technology
Figure 2. Schematic of the formation of 2,3,6-trichlorophenol radical and phenoxy radical from the reactions of 2,3,6-trichlorophenol with H. Potential barriers (∆E, kcal/mol) and the heats of reaction (∆H, kcal/mol) are shown.
26
ACS Paragon Plus Environment
Page 26 of 31
Page 27 of 31
Environmental Science & Technology
g=2.0035 48 h after heating
g=2.0042
298 K after heating 523 K 423 K 298 K
3320
3340
3360
3380
Figure 3. Electron spin resonance spectra of 2,3,6-trichlorophenol on the surface of silica-supported copper(II) oxide at (a) 298 K, (b) 423 K, (c) 523 K, and (d) 298 K after heating and (e) 48 h after heating.
27
ACS Paragon Plus Environment
Environmental Science & Technology
Page 28 of 31
523 K
523 K
g=2.0073
350 K g=2.0073 298 K 3300
3320
3340
3360
3380
Figure 4. Dynamic changes in the electron spin resonance spectra of 2,3,6-trichlorophenol on the surface of silica-supported copper(II) oxide at 298 K to 523 K. The inset shows a comparison of the experimental (green) and the simulated (red) spectra.
28
ACS Paragon Plus Environment
Page 29 of 31
Environmental Science & Technology
Cl
Cl
Pathway 1 Cl
Cl
Cl
Cl
-e-
-H2O Cl
Cl
Cl
Cl
Cl
OH
O HO
O
OH
Cl
OH
O
OH
Cu2+
Cu2+ Cl
Pathway 2 Cl
OH
HO
Cu+
O
Cu+
OH
Cu+ Cl
HO
OH
-H2O -HCl
Cl
Cl
O
-e-
Cl
Cl
O
O
Cu2+
Cl
O
O
OH
Cu+
Cl
O
O
Cu+
PCDD/Fs (cf. Figure 7)
Cu+
Cu2+ Cl
Cl
Cl Cl
Cl
-HCl
Pathway 3 OH
O
OH
O
OH
OH
Cu2+
Cu2+
Cl
OH
OH
OH
Cl HO
Cl
Cl
-e-
Cu+
O
OH
Cu+
Figure 5. Three possible pathways for the formation of persistent free radicals of 2,3,6-trichlorophenol radical from 2,3,6-trichlorophenol on the surface of silica-supported copper(II) oxide.39
29
ACS Paragon Plus Environment
Environmental Science & Technology
Cl
Page 30 of 31
Cl O
O
ring closure and intra-annular elimination of Cl Cl
O Cl
Cl
O
1,4,6,9-TCDD (1) Cl
Cl
Cl
Cl
Cl
O
Cl
Cl O
Cl
Cl
Cl
O
-Cl
1,2,6,9-TCDD (1)
O Cl
Cl
Cl
Cl
IM1
Cl
Cl
Cl
Cl
Cl O
O
O Cl
Cl
OH
Cl
Cl
Smiles rearrangement
Cl
Cl
Cl
Cl
O
ring closure and intra-annular Cl Cl elimination ofCl Cl
O
O Cl
Cl
1,2,6,9-TCDD (2) Cl
O
dimerization O
O
Cl
Cl Cl
O
Cl
Cl
O
Cl Cl
O
Cl
Cl
O
O
O Cl
Cl O Cl
Cl
Cl Cl Cl
Cl
dimerization
+
1,2,6,9-TCDD
ring closure and intra-annular elimination of Cl
Cl
Cl
O Cl Cl
Cl
1,4,6,9-TCDD (2)
Cl
Cl
Cl
Cl
IM2
O
Smiles rearrangement Cl
Cl
Cl
Cl
Cl
Cl
Cl
1,2,6,7-TCDD
O Cl
Cl
Cl
Cl
O
Cl
Cl Cl
Cl
O
O O O
dimerization
ring closure and intra-annular elimination Cl of Cl
Cl
Cl
Cl
Cl
Cl
Cl
O
Cl
Cl
O
O
O Cl
Cl Cl
O
O
Cl
1,2,6,9-TCDD Cl
O
Cl
Cl
Cl O
O
Cl
Cl
1,2,8,9-TCDD
O
O
ring closure and
Cl
Cl
Cl
Cl intra-annular
Cl
Cl
Cl
elimination of Cl
IM3
1,2,6,7-TCDD
Cl
O
Smiles rearrangement
O
1,2,6,9-TCDD
Cl Cl
Cl
Cl
Cl
Cl
Cl
Cl
Cl
Cl
Cl
Cl
O
O
O O O
Cl
Cl
Cl
Cl
Cl Cl
O
ring closure and intra-annular elimination ofClCl
Cl
O
1,2,6,9-TCDD Cl
Cl O
Cl
O Cl
O Cl Cl
1,2,6,7-TCDD
Figure 6. Proposed formation routes of tetrachlorodibenzo-p-dioxins (TeCDDs) involving the following five elementary processes: dimerization of 2,3,6trichlorophenol radicals, ortho-chloride abstraction, Smiles rearrangement, ring closure, and intra-annular elimination of chloride.
30
ACS Paragon Plus Environment
Page 31 of 31
Environmental Science & Technology
Graphic Abstract
31
ACS Paragon Plus Environment